Journal of Alloys and Compounds: Soraya Darafarin, Reza Sahraei, Ali Daneshfar

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Journal of Alloys and Compounds 658 (2016) 780e787

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


journal homepage: http://www.elsevier.com/locate/jalcom

Effect of deposition temperature on structural and optical properties


of chemically grown nanocrystalline Ni doped ZnS thin films
Soraya Darafarin, Reza Sahraei*, Ali Daneshfar
Department of Chemistry, University of Ilam, P.O. Box: 65315-516, Ilam, Iran

a r t i c l e i n f o a b s t r a c t

Article history: ZnS:Ni thin films with high purity in composition and crystallographic sense were deposited by chemical
Received 10 August 2015 bath deposition method at temperatures ranging from 40 to 80  C. The effect of deposition temperature
Received in revised form on the average crystal size, lattice constant, lattice strain, optical, and photoluminescence properties of
27 October 2015
the films is followed and discussed. The band gap energy of the ZnS:Ni thin films deposited at 40  C is
Accepted 29 October 2015
blue-shifted by about 0.14 eV with respect to the bulk value, while with increasing deposition temper-
Available online 9 November 2015
ature to 80  C it converges to 3.6 eV. The photoluminescence (PL) spectra of the ZnS:Ni thin films
exhibited three emission peaks, two intense peaks located at 430 nm and 523 nm and a weak shoulder
Keywords:
Nanostructured materials
peak over 325e400 nm range. The intensity of photoluminescence emission is enhanced with increasing
Semiconductors deposition temperature. X-ray diffraction analysis shows that the ZnS:Ni films constituted from nano-
Thin films crystals possessing average radii close to, and partially larger than, the radius of a bound exciton in the
Optical properties corresponding bulk semiconductor. Finally, a subsequent analysis based on effective mass approximation
model permits determination of the reduced effective mass of the electronehole pair ðm*eff Þ and the
relative dielectric constant (εr) for the ZnS:Ni nanocrystals in thin film from.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction new electronic states and can assist in the creation of new optical
centers to produce different colored emissions [14]. In the other
Zero-dimensional nanoscale materials such as quantum dots words, the Ni2þ ions induce new impurity states at different posi-
and semiconductor nanocrystals exhibit interesting and useful tions within the host band gap and strongly interfere with the
properties and may be applied in thin films form for the fabrication existing carrier recombination process [12]. As a consequence, the
of the next generation of nanoelectronic and nanooptic devices [1]. recombination process occurs via these newly induced electronic
Among the various nanocrystalline semiconductors, zinc sulfide states and results in different emission colors [18].
(ZnS) has attracted intensive studies due to its potential applica- The chemical bath deposition (CBD) method has been used for
tions in many optoelectronic and photovoltaic technologies such as the doping of various transition metals ions into the ZnS host
light emitting diodes, sensors, and thin film solar cells [2e6]. crystals in thin film form [19e27]. However, there are few reports
Pure ZnS crystals have large band gap energy of about 3.6 eV for the specific deposition of Ni doped ZnS thin films by the CBD
(cubic zinc blende phase) and therefore their applications are method [16,28e30]. Huang et al. [28] reported the preparation of
limited to blue and ultraviolet light emitting devices [7,8]. When the undoped and Ni doped ZnS thin films on ITO-coated glass
transition metals (dn) ions are doped into the ZnS host crystals, the substrates at 80  C using the sodium citrate as complexing agent in
optical, electrical, and magnetic properties of the host lattice are alkaline bath (pH > 11). They reported that doping of the ZnS thin
improved [9e13]. Recently, nickel doped zinc sulfide (ZnS:Ni) films with Ni leads to a considerable red shift in absorption edge of
nanocrystals have attracted more attention because of their tunable the UVevis spectra. The band gap values obtained by Huang et al.
fluorescence emission and low temperature ferromagnetic prop- for the undoped and Ni doped ZnS films were in the range of
erties [13e17]. It is known that, the Ni2þ ions as dopant can create 3.37e3.02 eV which were somewhat smaller than the typical value
of bulk ZnS (~3.6 eV). They have concluded that the decrease of the
band gap is probably related to the annealing effect and the
deformation of the films structure. In addition, they have investi-
* Corresponding author. gated and compared the photoelectrical properties of the undoped
E-mail addresses: r.sahraei@ilam.ac.ir, reza_sahrai@yahoo.com (R. Sahraei).

http://dx.doi.org/10.1016/j.jallcom.2015.10.272
0925-8388/© 2015 Elsevier B.V. All rights reserved.
S. Darafarin et al. / Journal of Alloys and Compounds 658 (2016) 780e787 781

and Ni doped ZnS thin films in details. Recently, Akhtar and co- excitation of the ZnS:Ni films in the UV region. All measurements
workers [29] have prepared the Ni doped ZnS thin films by CBD were carried out at room temperature.
method at 80  C and in acidic solution (pH ¼ 3.8). They have
studied the ferromagnetic and half metallic properties of the 3. Results and discussion
deposited films based on experimental results and theoretical cal-
culations in details. In our previous works, we have deposited the 3.1. Elemental analysis
ZnS:Ni thin films in a weak acidic bath (pH ¼ 6.0) at 80  C and the
effect of Ni doping on compositional, morphological, structural, and Chemical composition of ZnS:Ni thin films were analyzed by
optical properties of the films has been investigated in details EDX and ICP-AES measurements. EDX analysis identified the pres-
[16,30]. Nevertheless, to the best of our knowledge, no evidence of ence of zinc, sulphur, and silicon for all the layers deposited (Fig. 1).
the effect of deposition temperature on the structural and optical The signal of silicon comes from the silicon substrate. The Zn/S
properties of the Ni doped ZnS thin films prepared by the CBD atomic ratios for the prepared thin films were summarized in
method have been reported. Table 1. As can be seen, the variation of the Zn/S atomic ratios for all
In the present study, the Ni doped ZnS thin films were deposited thin films deposited at different temperatures is small and the Zn/S
by the CBD method at temperatures ranging from 40 to 80  C at pH atomic ratios are almost close to their stoichiometry. As the actual
of 6. The weak acidic bath (pH of 6.0) is chosen to avoid the for- concentration of Ni doped in all of the films was less than the
mation of undesired species such as ZnO and/or Zn(OH)2 in the detection limit of EDX, therefore, the Ni concentration in the ZnS:Ni
films [16,20,31,32]. The aim of this present work is to study the films was determined by ICP-AES measurements. The obtained Ni/
effect of deposition temperature on the structural, optical and Zn atomic percent ratios for the ZnS:Ni films deposited at 40, 60
photoluminescence properties of the ZnS:Ni thin films. The films and 80  C for 8 h were also listed in Table 1. ICP-AES analysis shows,
are characterized by several techniques such as X-ray diffraction, with increasing in deposition temperature, the Ni/Zn atomic ratios
field emission scanning electron microscopy (FE-SEM), optical in the films increase. It should be noted that a very weak peak (less
transmittance and fluorescence spectrophotometer. Measure- than the detection limit of EDX) at around 1 KeV in the EDX spec-
ments, such as those outlined above, allow quantitative analysis of trum (Fig. 1) could be corresponding to trace amounts of oxygen in
the increase in band gap energy due to quantum confinement ef- the films. According to our previous works, this very small amount
fects between the conduction and valence bands. A subsequent of oxygen is related to adsorbed water on the films as it has been
analysis of these data permits determination of the reduced detected by FT-IR and XPS analysis [16,20,31,32]. The co-deposition
effective mass of the electronehole pair and the relative dielectric of ZnS and NiS quite difficult, because of the ZnS (Ksp ¼ 1.1  1024)
constant for the ZnS:Ni nanocryatals in thin films from. These is much less soluble than the NiS (Ksp ¼ 1.1  1020). During the
values are compared, where available, with previously determined deposition process, ZnS can forms the main phase, whereas Ni2þ
values for the bulk ZnS. ions are present as an impurity and can be incorporated gradually
into the ZnS crystal lattice, depending on the deposition
2. Experimental temperature.

The ZnS:Ni films were deposited at different temperature on


3.2. Structural and morphological characterization
quartz, glass and silicon substrates from aqueous solutions of 6 mL
of a 1 M (M ¼ molL1) zinc acetate solution, 15 mL of a 0.2 M
The X-ray diffraction (XRD) patterns of the ZnS:Ni thin films
Na2EDTA solution, and nickel acetate solution was then added in
prepared at different deposition temperatures was shown in Fig. 2.
desired amounts. The Ni2þ to Zn2þ molar ratio in the precursor
The line XRD patterns correspond to bulk cubic ZnS (Joint Com-
solution was 1:10. Then 30 mL of a 0.4 M thioacetamide solution
mittee for Powder Diffraction Standards, JCPDS card No. 05-0566)
was added and the pH was adjusted to 6.0 by adding a 1 M acetic
and bulk rembohydral NiS (JCPDS card No. 12-0041) were given at
acid solution. Double distilled water was added to the solution to
the bottom of Fig. 2. As can be seen, the XRD patterns for nano-
make the volume close to 100 mL and pH was adjusted to 6.0 again.
crystalline Ni doped ZnS thin films exhibit main diffraction features
The procedure for cleaning substrates was described in our previ-
ous work [16]. This solution was poured into a glass tank (reaction
bath) for the film deposition. Substrates were immersed vertically
in the glass tank and the glass tank was placed in a thermostat bath
at the desired temperature. The deposition was carried out at 40,
60, and 80  C for 8 h, respectively. The deposited films were then
washed with distilled water and dried in air at room temperature.
The atomic composition of the films was analyzed by energy-
dispersive X-ray spectrometer (EDX) using an Oxford INCA II en-
ergy solid state detector. Inductively coupled plasma atomic
emission spectroscopy (ICPAES; Varian Vista-Pro) was employed to
determine the elemental content of the ZnS:Ni thin films. X-ray
diffraction (XRD) patterns were recorded with an automated Phi-
lips X'Pert X-ray diffractometer with Cu Ka radiation (40 kV and
30 mA) for 2q values over 20e60 . The surface morphology of the
ZnS:Ni films were observed by field emission scanning electron
microscopy (FE-SEM; Hitachi S-4200) under an acceleration voltage
of 15 kV. The UVevis spectra were recorded by a double beam
Perkin Elmer Lambda 25 spectrophotometer in the spectral range
from 300 to 800 nm. The photoluminescence emission and exci-
tation spectra of the films were measured using a Cary Eclipse Fig. 1. Typical EDX spectrum of the ZnS:Ni thin film deposited on silicon substrate at
fluorescence spectrophotometer. A xenon lamp was used for 80  C for 8 h.
782 S. Darafarin et al. / Journal of Alloys and Compounds 658 (2016) 780e787

Table 1
The film thickness, t, Zn/S and Ni/Zn ratios, the lattice constant, a, the lattice strain, h, the average crystal size, D, and the band gap energy, Eg, values for the ZnS:Ni films
deposited at the different temperatures.

Temp.( C) t (nm) Zn/S (at ratio) Ni/Zn (at %) a (nm) h D (nm) Eg (eV)

40 170 1.02 0.14 0.5383 0.0072 4.30 3.74


60 280 0.99 0.20 0.5391 0.0067 5.9 3.70
80 430 1.01 0.26 0.5405 0.0060 9 3.63

On the basis of X-ray diffraction data, it is quite easy to calculate


the lattice constant (a) for cubic structure from the d-interplanar
spacing of all observed peaks in XRD patterns using the following
equations [35]:

. 1=2
dhkl ¼ a h2 þ k2 þ l2 and nl ¼ 2dhkl sin qhkl (1)

where dhkl is the d-interplanar spacing with Miller indices h, k, and


l, qhkl is the corresponding diffraction Bragg angle, while l is X-ray
wavelength of the CuKa1 radiation, and n is the diffraction order.
For n ¼ 1, with combining of the two above equations we could
easily derived the following expression:
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
l h2 þ k2 þ l2
a¼ (2)
2 sin qhkl

On the basis of data for all peaks in the X-ray diffractograms of the
ZnS:Ni thin films prepared at different deposition temperatures,
the corresponding values of the lattice constant (a) were calculated.
Fig. 2. XRD patterns of the ZnS:Ni films deposited at different deposition Then, in order to avoid from random and systematic errors on
temperatures. the measurement of qhkl, the result is plotted against
FðqÞ ¼ cos2 q=sin q þ cos2 q=q (Fig. 4). By vertical intercept of the
straight-line fit of diagrams in the Fig. 4, the precise values of the
correspond to (111), (220), and (311) planes; the peaks position lattice constant (a) for the ZnS:Ni thin films deposited at different
match well with the standard pattern of the ZnS cubic phase [32]. temperatures are obtained [32,36]. The results are given in Table 1.
The diffraction peaks in all diffractograms are relatively broadening As can be seen, the lattice constants for the ZnS:Ni thin films
due to finite crystalline size of the ZnS:Ni thin films [33]. It is worth deposited at different temperatures are slightly smaller than that
noting that, no diffraction peaks correspond to NiS has been reported for the bulk ZnS cubic structure (0.5406 nm), and it con-
detected, indicates that the Ni2þ ions preferably incorporated into verges to the bulk value upon increasing deposition temperature.
the ZnS lattice and there is no secondary phase in the structure of As can be seen in Fig 2, the X-ray diffraction patterns for all thin
thin films. Also, no characteristic peaks related to the impurity films deposited at different deposition temperature are relatively
phases such as ZnO and/or Zn(OH)2 were detected in the ZnS:Ni broad. Tow important factor play basic role on the broadening of X-
thin films. It is clearly visible from Fig. 2, for the deposited film at ray diffraction peaks. One of these factors is the approximate crystal
80  C, the intensity of diffraction peaks are relatively stronger than diameter (D) and another is the lattice strain (h). In order to eval-
that of the films deposited at low temperatures. This suggests that uate the contribution of each of these two factors, we start by
the following factors are contributed in such behavior for the thin expressing the overall broadening (b) of the XRD peaks by the
films [34]: following relation [37,38]:

i the deposited films at high temperature have more thickness. In b cos q ¼ 4h sin q þ Kl=D (3)
fact, the films with high thickness show the diffraction peaks
with high intensity in the XRD patterns. where, K is the shape factor in spherical approximation which is
ii at high deposition temperature, the bigger nanocrystals is approximately unity, l is the wavelength of the used X-ray radia-
grown. This temperature dependence of nanocrystals size may tion, q is the angle correspond to the diffraction maximum, and b is
be explained on thermodynamic grounds of larger crystallites the full-width at half-maximum (FWHM) of the XRD peaks
being more stable than the smaller ones which are more favored appearing at the diffraction angle q. The b parameter in the above
kinetically at lower temperatures. equation must be corrected with the instrumental width through
iii at high deposition temperature, the Ni2þ ions are more incor- using of the geometric mean [39]:
porated into the vacancy states of the host matrix, therefore, the
degree of crystallinity of the ZnS:Ni thin films is improved.  1=2
  1=2
b¼ bexp  binst b2exp  b2inst (4)
The cross-sectional view of SEM images for the ZnS:Ni thin films
deposited on glass substrate at deposition temperatures of 40 and
where, bexp and binst parameters are the experimental and the
80  C was shown in Fig. 3(a and b), respectively. It can be also seen
instrumental linear widths in radians, respectively. The value of
that the thickness of the ZnS:Ni thin films is increased with
binst was determined to be 2.58  103 rad by using a standard
increasing deposition temperature.
silicon powder. Thus, it is clear that when bcosq is plotted against
S. Darafarin et al. / Journal of Alloys and Compounds 658 (2016) 780e787 783

Fig. 3. Cross-sectional SEM images of the ZnS:Ni thin films deposited at 40  C (a) and at 80  C (b).

films are calculated, respectively. All of these parameters relevant


to the present study are given in Table 1, together with the relevant
optical spectroscopy data obtained at room temperature. Briefly, it
could be concluded from Table 1 that the lattice strain (h) values in
the case of all ZnS:Ni thin films deposited at different temperatures
are notably small and, in all cases, they decrease with increasing
deposition temperature. The similar results were also reported for
nanostructured ZnSe and CdSe thin films by Pejova et al. [40,41].
These small values of the lattice strain indicates that the nano-
crystals size factor plays important role on the broadening of X-ray
diffraction peaks of the ZnS:Ni thin films deposited at different
temperatures. Moreover, the thin films deposited by CBD are under
very low tensile strain along the (111) plane parallel to the substrate
surface. It can be seen that as the deposition temperature increased
from 40  C to 80  C, the nanocrystales diameter was slightly
increased from 4.30 nm to 9 nm. In summary, these results show
that the size of ZnS:Ni nanocrystals depends mainly on the depo-
sition temperature. Furthermore, because of very small size of the
ZnS:Ni nanocrystals in the thin films, the interplanar distance, d111,
of the planes parallel to the films surface are smaller than the
Fig. 4. Plots of the lattice constant, a, of the ZnS:Ni films against
FðqÞ ¼ cos2 q=sin q þ cos2 q=q. The lattice constants were calculated from different XRD corresponding bulk literature value (0.31260 nm) and it converges
peaks. to the bulk value upon deposition temperature.
Fig. 6(a and b) show the plan-view of the SEM images of thin
films deposited on glass substrate at different temperatures of 40
sinq a straight line with slope 4h and intercept Kl/D is obtained, as and 80  C. As shown in images, the film surface became relatively
shown in Fig. 5. From the slope and intercept of these straight lines, compact and uniform throughout all the regions, without the
the lattice strain and the average nanocrystals diameter of the thin presence of voids, pinholes, or cracks at all temperatures. At 40  C
the surface of film was covered with small grains about 40e50 nm
in size. When the temperature reaches to 80  C, the film consists of
grains and reveals some size categories of 50, 100, and 150 nm, also
larger clusters of about 250 nm in size are formed. It can be seen
that the clusters themselves are aggregation of small grains. As a
result, the deposition temperature has an significant effect on the
grains size of the ZnS:Ni thin films. It is to note that the XRD
analysis reveals the fact that the grains themselves must be coa-
lescence of nanocrystals of about 4e9 nm in size as given in Table 1.
This suggests that each feature in the SEM images is a poly-
crystalline grain.

3.3. Optical properties

The room temperature photoluminescence (PL) behavior of


uncoated quartz slide as a blank as well as that the Ni doped ZnS
thin films grown on quartz substrates at 60  C were investigated
using fluorescence spectrophotometer and the results are shown in
Fig. 7 (a). The excitation wavelength was 245 nm. It is important to
note that in the most of previous reports on the ZnS:Ni nano-
Fig. 5. Plots of bcosq versus sinq for the ZnS:Ni films deposited at different deposition structures only one peak at 520 nm was observed [42,43]. But, the
temperatures. PL spectrum of the ZnS:Ni films consists of three characteristic
784 S. Darafarin et al. / Journal of Alloys and Compounds 658 (2016) 780e787

Fig. 6. SEM images of the ZnS:Ni thin films deposited at 40  C (a) and at 80  C (b).

Fig. 7. (a) PL spectra of the uncoated quartz and the ZnS:Ni film deposited on quartz substrate at 60  C for 8 h, and (b) PL spectra of the ZnS:Ni films deposited at various
temperatures.

bands, two intense bands located at 430 nm and 523 nm and a While the PL emission intensity for the ZnS:Ni films deposited at 60
weak shoulder peak over 325e400 nm range. It is well known that and 80  C was enhanced. The high PL intensity could be related to
the blue emission at 430 nm is related to the defect levels within more incorporation of the Niþ2 ions into the ZnS host crystals as
the band gap, where interstitial sulfur and vacancy states were well as improvement in crystallinity and thickness of the ZnS:Ni
formed during the deposition of nanocrystalline ZnS films. The films deposited at higher temperature [47]. Hence, the more new
similar results were also reported for the blue emission of ZnS radiative centers were generated in the ZnS host crystals at high
nanopraticles in the literature [44,45]. The green emission at deposition temperature. Thus, the decrease of defects density and
523 nm of the ZnS:Ni films can be attributed to the ded optical the increase of radiation centers in the ZnS:Ni films deposited at
transitions of Ni2þ luminescent centers formed in the ZnS host higher temperature play an important role in the improvement of
crystals. When Ni2þ ions incorporate into the ZnS lattice, the lowest their optical and photoluminescence properties. Therefore, the
multiplet term 3F of the free Ni2þ ions is split into 3T1, 3T2, and 3A2 nanocrystalline ZnS:Ni thin films with intense green emission may
through the anisotropic hybridization [11,46]. In summary, the have the potential applications in electroluminescent devices such
photoluminescence in this region is attributed to the 3T2 / 3A2 as large-area displays.
transitions within the 3d shell of Ni2þ ions that incorporated into The optical transmission and reflection spectra of bare glass and
the ZnS lattice. In addition, the weak shoulder UV emission peak in the ZnS:Ni thin films deposited on glass substrates at 40, 60, and
the range of 325e400 nm corresponding to near-band-edge 80  C for 8 h were studied in wavelengths region from 300 to
emission is originated from the recombination of free excitons of 800 nm and shown in Fig. 8 (a). As can be seen, all thin films have
ZnS [18]. high transparency in the visible region. The optical transmittance of
Fig. 7(b) shows the PL emission spectra of the ZnS:Ni films the ZnS:Ni thin films was varied between 45 and 90% in the visible
prepared at different deposition temperatures for 8 h. It well region and was depended on the deposition temperature. The
known that the PL intensity of semiconductors was closely related decrease in the optical transmission of thin films with increasing
to two important factors: the quality of crystals and the abundance temperature can due to enhancement in the films thickness as can
of radiative centers in the films. As shown in Fig. 7(b), the ZnS:Ni be seen from the cross sectional FE-SEM images (Fig. 3(b)).
film deposited at 40  C has the lowest PL intensity. The low in- The absorption edge of the ZnS:Ni thin films is observed at about
tensity of the PL emission was suggested that the crystallinity of the 330e340 nm. The sharp absorption edge is due to the uniform and
ZnS:Ni films was low (as seen in the XRD Pattern), and also the Ni2þ compact morphology of the ZnS:Ni thin films [18]. It can be seen
ions were not incorporated effectively into the ZnS host crystals. In that the absorption edge of the spectra clearly shifted towards
addition, the low thickness (170 nm) of the ZnS:Ni films deposited shorter wavelengths in the lower deposition temperature. This
at 40  C could be another reason for their low PL emission intensity. blue-shift in absorption edge is associated with increasing in the
S. Darafarin et al. / Journal of Alloys and Compounds 658 (2016) 780e787 785

Fig. 8. Optical transmittance and reflectance spectra (a), and dependence of absorption coefficients on incident photon energy (b) of the ZnS:Ni thin films deposited at various
temperatures.

band gap energy of the ZnS:Ni films. Based on the obtained optical
absorbance data, the values of absorption coefficient, a, could be
determined by using the following equation [48,49]:

A ¼ lnðI0 =It Þ ¼ ad (5)

where I0, It are incident and transmitted light intensities, respec-


tively, d is thickness of the film, and A is the optical absorbance. The
dependencies of the absorption coefficients versus incident pho-
tons energy (hn) for the nanocrystalline ZnS:Ni thin films deposited
at different temperatures were shown in Fig. 8 (b). As can be seen,
The magnitude of absorption coefficients for the ZnS:Ni thin films
in the absorption edge region are of the order of 105 (cm1) and
these suggest that the interband electronic transitions are of the
allowed direct type. In the absorption edge region, the electronic
transitions were produced free electrons and holes in the conduc-
tion and valence bands, respectively. In addition, based on the band
theory, it can be concluded that the fundamental absorption of all
thin films deposited at different temperatures correspond to direct
electrons excitation from the highest valence band to the lowest Fig. 9. Plots of (ahy)2 versus hy for the ZnS:Ni thin films as a function of the different
conduction band. deposition temperatures.
When the interband electronic transitions are the allowed
direct, then we can determine the band gap energy (Eg) of the
ZnS:Ni thin films deposited at different temperatures based on the nanocrystals. This effect should be significant when the radius of
fitting of absorption coefficients data to the following equation nanocrystals becomes comparable to or smaller than the Bohr
[44]: radius of the exciton (rB) for the bulk corresponding semiconductor.
The rB is given by the following relation [51,52]:
 
ðahnÞ2 ¼ K hn  Eg (8) ε0 εr h2
rB ¼ (9)
pe2 m*eff
where K is a constant which arises from the Fermi's golden rule for
fundamental interband electronic transitions, while all other
symbols have the same meaning as in the previous equations. The where h is the Planck constant, e is the charge of electron, ε0 is the
band gap energy, Eg, of thin films can be calculated by plotting permittivity of vacuum, εr is the relative dielectric constant of the
(ahy)2 versus hy and then extrapolating the straight-line portion of bulk semiconductor, and m*eff is the reduced effective mass of the
the curve to zero absorption coefficients (a ¼ 0). As shown in Fig. 9, electronehole pair given by the following equation [53].
the band gap of the films varies from 3.63 to 3.74 eV with the
variation of deposition temperature where the deposition time is m*e m*h
m*eff ¼ (10)
same for all the films. The obtained band gap values are listed in m*e þ m*h
Table 1. These values are slightly larger than the typical value of the
bulk ZnS (3.6 eV). These blue-shift in the band gap energy of our In Eq. (10), m*e and m*h are the effective masses of the electron and
chemically deposited ZnS:Ni thin films, compared with the corre- hole, which are expressed in unit of the free electron mass (m0).
sponding value of bulk semiconductor, is due to the quantum Using the corresponding parameters for ZnS, the estimated value
confinement effects in nano-size particles [50]. for rB is about 3.2 nm. The estimated values of the rB for ZnS in
The quantum size effect plays an important role in determining previous literature is approximately 2.8 nm [54,55]. In our samples,
the electronic and optical properties of semiconducting depending upon the deposition temperature, the average
786 S. Darafarin et al. / Journal of Alloys and Compounds 658 (2016) 780e787

nanocrystals radius found to be about 2.15, 2.95, and 4.5 nm based


on XRD results as given in Table 1. As can be seen, the nanocrystals
radius in our thin films is quite in line with the exciton Bohr radius
(nanocrystals radius ~ exciton radius) and hence the quantum
confinement effects should be significant.
As the nanocrystals size is in the range between the sizes of
molecules and bulk semiconductors, the quantum size effect can be
explained with hybrid molecular and semiconductor languages.
The linear combination of atomic and molecular orbitals provides a
natural framework to understand the evolution of nanocrystals
from molecules to bulk and the size dependence of the lowest
excited-state energy (band gap). In the effective mass approxima-
tion (Brus model), with contribution of the Coulomb attraction and
the spatial correlations between the two particles, the approximate
solution to the Schro€dinger wave equation for an electron and hole
in a sphere has been used to model the electronic properties of
semiconductor nanocrystals. The calculations provide an analytical
expression for how the electronic band gap of the semiconductor
nanocrystal (EgNC ) is modified relative to that of the bulk semi-
conductor (EgB ). In other words, the energy of quantum confinement
Fig. 10. Plot of RDE against R1for the ZnS:Ni thin films.
(DE ¼ EgNC  EgB ) is directly related to the nanocrystal radius (R)
[56,57]:
electronehole pair and the only variable in this term is the relative
dielectric constant (εr). In the effective mass approximation model,
h2 1:8e2 e4 m*eff
DE ¼   0:248 (11) the Coulomb term in Eq. (12) is usually calculated by using the bulk
8m*eff R2 4pε0 εr R 2ðε0 εr Þ2 h2 value for εr. There are, however, strong indications in many litera-
ture related to the quantum size effects that the dielectric screening
The first term in Eq (11) represents the particle-in-a-box quantum should be decreased for small-sized semiconductor nanocrystals
localization (i.e., the kinetic energy term), and it shifts the DE to [51,57,61]. Such a decrease in the dielectric screening could be due
lower energies proportionally to 1/R2. The second term is due to the to the inability of the lattice polarization to follow the more rapid
screened Coulomb interaction between the electron and the hole, electron and hole motions associated with a smaller crystal radius.
and has simple 1/R dependence. The last size-independent term in
Eq. (11) is a result of the spatial correlation between the two par- 4. Conclusions
ticles and is usually small in both the absolute value and also in the
sense of its contribution to the overall DE for wide band gap In conclusion, we have developed a simple method to produce
semiconductors with relatively large dielectric constant [51,56]. In the Ni doped ZnS nanocrystals in thin film form at different
the our present case, the ZnS:Ni nanocrystals have a wide band gap deposition temperatures with potential applications in optoelec-
in the range of 3.6e3.8 eV with a large relative dielectric constant of tronics and solar cell engineering. No secondary phase, apart from
about 8.6 and therefore the spatial correlation contribution in the cubic lattice structure of the ZnS, was observed in the XRD patterns.
size-dependent shift of band gap energy (DE) could be neglected. SEM images indicated that the surface morphology of the ZnS:Ni
Accordingly, by ignoring of the third term in Eq. (11), we can easily films strongly depends on deposition temperature. The green
rearrange this equation as follows: emission at 523 nm of the ZnS:Ni films can be related to the ded
optical transitions of Ni2þ luminescent centers formed in the ZnS
  host crystals. Furthermore, with increase in the deposition tem-
h2 1 1:8e2
RDE ¼  (12) perature, the PL emission intensity of the ZnS:Ni films is improved
8m*eff R 4pε0 εr
that can be attributed to the decrease of crystal defects density and
also an enhancement in number of the radiation centers. The band
It is clear that the points in a plot of (RDE) against (1/R) are con- gap energy is decreased with increasing deposition temperature
nected by a straight line with slope ðh2 =8m*eff Þ and intercept and it is accompanied with an increase in the average crystal size of
ð1:8e2 =4pε0 εr Þ, respectively; as shown in Fig. 10. From the slope the ZnS:Ni films, lattice strain relaxation, and lattice constant
and intercept of this straight line, the reduced effective mass of the enlargement. Also, based on effective mass approximation model,
electronehole pair ðm*eff Þ and the relative dielectric constant (εr) for the values of the reduced effective mass of the electronehole pair
the ZnS:Ni nanocrystals are calculated, respectively. The resulting ðm*eff Þ and the relative dielectric constant (εr) for the nanocrystalline
values for m*eff and εr, being 0.18m0 and 6.5, respectively, are ZnS:Ni thin films were determined to be 0.18m0 and 6.5,
respectively.
somewhat smaller than those reported for the bulk ZnS (i.e. m*eff ¼
0:28m0 and εr ¼ 8:6); probably due to the quantum size effect as References
expected from the nanocrystalline nature of the ZnS:Ni thin films
[58e60]. In order to explanation of the decrease in the m*eff and εr [1] Y.W. Jun, J.S. Choi, J. Cheon, Angew. Chem. Int. Ed. 45 (2006) 3414e3439.
[2] X. Yuan, J. Hua, R. Zeng, D. Zhu, W. Ji, P. Jing, X. Meng, J. Zhao, H. Li, Nano-
with respect to their bulk values, we should be consider an technology 43 (2014) 435202.
important effect that leading to the inadequacy of the Brus model, [3] A.M. Diamond, L. Corbellini, K.R. Balasubramaniam, S. Chen, S. Wang,
which is usually called the size-dependence of the Coulomb T.S. Matthews, L.W. Wang, R. Ramesh, J.W. Ager, Phys. Status Solidi A 209
(2012) 2101e2107.
interaction. As mentioned before, the second term in Eq. (12) is [4] X. Fang, T. Zhai, U.K. Gautam, L. Li, L. Wu, Y. Bando, D. Golberg, Prog. Mater. Sci.
related to the screened Coulomb interaction between the 56 (2011) 175e287.
S. Darafarin et al. / Journal of Alloys and Compounds 658 (2016) 780e787 787

[5] J. Kim, C. Park, S.M. Pawar, A.I. Inamdar, Y. Jo, J. Han, J.P. Hong, Y.S. Park, [32] R. Sahraei, G. Motedayen Aval, A. Goudarzi, J. Alloys Compd. 466 (2008)
D.Y. Kim, W. Jung, H. Kim, H. Im, Thin Solid Films 566 (2014) 88e92. 488e492.
[6] M. Nguyen, K. Ernits, K.F. Tai, C.F. Ng, S.S. Pramana, W.A. Sasangka, [33] H. Hennayaka, H.S. Lee, Thin Solid Films 548 (2013) 86e90.
S.K. Batabyal, T. Holopainen, D. Meissner, A. Neisser, L.H. Wong, Sol. Energy [34] R. Sahraei, G. Motedayen Aval, A. Baghizadeh, M. Lamehi-Rachti, A. Goudarzi,
111 (2015) 344e349. M.H. Majles Ara, Mater. Lett. 62 (2008) 4345e4347.
[7] J. Chen, F. Xin, S. Qin, X. Yin, Chem. Eng. J. 230 (2013) 506e512. [35] M.T. Weller, Inorganic Materials Chemistry, Oxford University Press, Oxford,
[8] A. Inamdar, S. Lee, D. Kim, K.V. Gurav, J.H. Kim, H. Im, W. Jung, H. Kim, Thin 1997.
Solid Films 537 (2013) 36e41. [36] A.E. Rakhshani, A.S. Al-Azab, Appl. Phys. A 37 (2001) 631e636.
[9] D.E. Ortíz-Ramos, L.A. Gonza lez, R. Ramirez-Bon, Mater. Lett. 124 (2014) [37] E. Lifshin, X-ray Characterization of Materials, Wiley-VCH, New York, 1999.
267e270. [38] B. Pejova, I. Bineva, J. Phys. Chem. C 117 (2013) 7303e7314.
[10] B. Poornaprakash, D.A. Reddy, G. Murali, N.M. Rao, R.P. Vijayalakshmi, [39] C. Suryanarayana, M.G. Norton, X-Ray Diffraction, Plenum Press, New York,
B.K. Reddy, J. Alloys Compd. 577 (2013) 79e85. 1998.
[11] M. Molaei, J. Lumin. 136 (2013) 38e41. [40] B. Pejova, I. Bineva, J. Mater. Sci. Mater. Electron. 26 (2015) 4944e4955.
[12] S. Jana, G. Manna, B.B. Srivastava, N. Pradhan, Small 9 (2013) 3753e3758. [41] B. Pejova, B. Abay, J. Phys. Chem. C 115 (2011) 23241e23255.
[13] J. Kaur, M. Sharma, O.P. Pandey, Opt. Mater. 47 (2015) 7e17. [42] Q. Pan, D. Yang, Y. Zhao, Z. Ma, G. Dong, J. Qiu, J. Alloys Compd. 579 (2013)
[14] D.A. Reddy, D.H. Kim, S.J. Rhee, C.U. Jung, B.W. Lee, C. Liu, J. Alloys Compd. 588 300e304.
(2014) 596e604. [43] V. Ramasamy, K. Praba, G. Murugadoss, Superlattices Microstruct. 51 (2012)
[15] S. Kumar, N.K. Verma, J. Mater. Sci. Mater. Electron. 25 (2014) 1132e1137. 699e714.
[16] R. Sahraei, S. Darafarin, J. Lumin. 149 (2014) 170e175. [44] D. Denzler, M. Olschewski, K. Sattler, J. Appl. Phys. 84 (1998) 2841e2845.
[17] S. Kumar, C.L. Chen, C.L. Dong, Y.K. Ho, J.F. Lee, T.S. Chan, R. Thangavel, [45] S. Wageh, Z.S. Linga, X. Xu-Rong, J. Cryst. Growth 255 (2003) 332e337.
T.K. Chen, B.H. Mok, S.M. Rao, M.K. Wu, J. Alloys Compd. 554 (2013) 357e362. [46] P. Yang, M. Lu, D. Xu, D. Yuan, J. Chang, G. Zhou, M. Pan, Appl. Phys. A 74
[18] S. Jana, B.B. Srivastava, S. Jana, R. Bose, N. Pradhan, J. Phys. Chem. Lett. 3 (2012) (2002) 257e259.
2535e2540. [47] N. Eryong, L. Donglai, Z. Yunsen, B. Xue, Y. Liang, J. Yong, J. Zhifeng,
[19] M.S. Akhtar, Y.G. Alghamdi, M.A. Malik, R.M.A. Khalild, S. Riaz, S. Naseem, S. Xiaosong, Appl. Surf. Sci. 257 (2011) 8762e8766.
J. Mater. Chem. C 3 (2015) 6755e6763. [48] J.I. pankove, Optical Processes in Semiconductors, Dover Pub., New York,
[20] A. Goudarzi, G. Motedayen Aval, S.S. Park, M.C. Choi, R. Sahraei, M. Habib 1967.
Ullah, A. Avane, C.S. Ha, Chem. Mater. 21 (2009) 2375e2385. [49] P. Roy, J.R. Ota, S.K. Srivastava, Thin Solid Films 515 (2006) 1912e1917.
[21] D.E. Ortíz-Ramos, L.A. Gonz alez, R. Ramirez-Bon, Mater. Lett. 124 (2014) [50] M.G. Bawendi, M.L. Steigerwald, L.E. Brus, Annu. Rev. Phys. Chem. 41 (1990)
267e270. 477e496.
[22] A.I. Inamdar, S. Lee, D. Kim, K.V. Gurav, J.H. Kim, H. Im, W. Jung, H. Kim, Thin [51] Y. Wang, N. Herron, J. Phys. Chem. 95 (1991) 525e532.
Solid Films 537 (2013) 36e41. [52] A. Mark Fox, Optical Properties of Solids, Oxford University Press, Oxford,
[23] A.C. Dhanya, K. Deepa, T.L. Remadevi, J. Mater. Sci. Mater. Electron. 24 (2013) 2001.
4782e4789. [53] B. Enright, D. Fitzmaurice, J. Phys. Chem. 100 (1996) 1027e1035.
[24] S. Muthukumaran, M.A. Kumar, Mater. Lett. 93 (2013) 223e225. [54] N. Chestnoy, R. Hull, L.E. Brus, J. Chem. Phys. 85 (1986) 2237e2242.
[25] H.J. Zhu, Y. Liang, X.Y. Gao, R.F. Guo, Q.M. Ji, Braz. J. Phys. 45 (2015) 308e313. [55] P.Y. Yu, M. Cardona, Fundamentals of Semiconductors: Physics and Materials
[26] Y. Chen, G.F. Huang, W.Q. Huang, B.S. Zou, A.L. Pan, Appl. Phys. A 108 (2012) Properties, fourth ed., Springer, Berlin, 2010.
895e900. [56] L. Brus, J. Phys. Chem. 90 (1986) 2555e2560.
[27] M.S. Akhtar, M.A. Malik, Y.G. Alghamdi, K.S. Ahmad, S. Riaz, S. Naseem, Mater. [57] B. Pejova, I. Grozdanov, Mater. Chem. Phys. 90 (2005) 35e46.
Sci. Semicond. process. 39 (2015) 283e291. [58] A.D. Yoffe, Adv. Phys. 42 (1993) 173e266.
[28] C.M. Huang, L.C. Chen, G.T. Pan, T.C.K. Yang, W.S. Chang, K.W. Cheng, Mater. [59] D.R. Vij, N. Singh, Luminescence and Related Properties of II-VI Semi-
Chem. Phys. 117 (2009) 156e162. conductors, Nova Science Pub. Inc, Commack, New York, 1998.
[29] M.S. Akhtar, M.A. Malik, S. Riaz, S. Naseem, Mater. Chem. Phys. 160 (2015) [60] D.R. Vij, Handbook of Electroluminescent Materials, IOP Publishing Ltd, Bris-
440e446. tol, 2004.
[30] R. Sahraei, S. Darafarin, Spectrochim. Acta Part A 149 (2015) 941e948. [61] Y. Wang, A. Suna, W. Mahler, R. Kasowski, J. Chem. Phys. 87 (1987)
[31] R. Sahraei, A. Daneshfar, A. Goudarzi, S. Abbasi, M.H. Majles Ara, F. Rahimi, 7315e7322.
J. Mater. Sci. Mater. Electron 24 (2013) 260e266.

You might also like